Sie sind auf Seite 1von 6

Chemical Engineering Science 61 (2006) 7133 7138 www.elsevier.

com/locate/ces

Experimental and numerical heat transfer in a plate heat exchanger


Flavio C.C. Galeazzo a , Raquel Y. Miura b , Jorge A.W. Gut c , Carmen C. Tadini c ,
a School of Chemical Engineering, State University of Campinas, Brazil b Escola Superior de Qumica, Faculdades Oswaldo Cruz, Brazil c Department of Chemical Engineering, Escola Politcnica, University of So Paulo, P.O. Box 61548, So Paulo, SP 05424-970, Brazil

Received 21 February 2006; received in revised form 12 July 2006; accepted 14 July 2006 Available online 22 July 2006

Abstract In the present study a virtual prototype of a four-channel plate heat exchanger with at plates was developed using computational uid dynamics (CFD). Parallel and series ow arrangements were tested and experimental results were compared to numerical predictions for heat load obtained from the 3D CFD model and also from a 1D plug-ow model. The CFD model represents channels, plates and conduits of the exchanger and takes into account the unequal ow distribution among channels and the ow maldistribution inside the channel. CFD results are in good agreement with experimental data, especially for the series arrangement. 2006 Elsevier Ltd. All rights reserved.
Keywords: Plate heat exchanger; Heat transfer; Mathematical modeling; Fluid mechanics; Computation; Food processing

1. Introduction Plate heat exchangers (PHEs) are extensively used for heating, cooling and heat-regeneration applications in the chemical, food and pharmaceutical industries because of their high thermal efciency, exibility and ease of sanitation. The PHE consists of a pack of metal plates pressed together into a frame. A sequence of thin channels is formed between the plates and the ow distribution of the hot and cold streams is dened by the plate perforations and the gasket designs; thus a large number of congurations is possible (Gut and Pinto, 2004). The thermal-hydraulic model of a PHE usually assumes (1) 1D plug-ow inside the channels and (2) uniform ow distribution throughout the exchanger with equal ow in all channels (Zaleski, 1984; Gut and Pinto, 2003). However, these assumptions are unrealistic and may not hold in practice (Mehrabian et al., 2000; Tereda et al., 2005). According to Rao et al. (2002) and Tereda et al. (2005), a PHE model based on the unequal ow distribution among channels is required for proper equipment evaluation and design.
Corresponding author. Tel.: +55 11 30912258; fax: +55 11 30912255.

E-mail address: catadini@usp.br (C.C. Tadini). 0009-2509/$ - see front matter 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2006.07.029

For modeling the ow distribution from channel to channel, Bassiouny and Martin (1984), Thonon and Mercier (1996) and Rao et al. (2002) analyzed the pressure distribution in the conduits of the PHE and determined the ow rate for each channel. The U-type and Z-type parallel pass arrangements were compared for analyzing the effect of unequal ow distribution among channels. In order to accurately model ow and heat transfer inside the channels of a PHE, 3D computational uid dynamics (CFD) methods need to be used. CFD is a simulation tool, which makes use of computational algorithms to model heat, mass and momentum transfer. In the recent years, CFD use in the food industry has been increasing, including PHE applications such as milk pasteurization where fouling plays a very important role (Xia and Sun, 2002; Fernandes et al., 2005). Kho and Mller-Steinhagen (1999) investigated the effect of ow distribution inside the channel of a PHE with at plates. Different distributors were tested experimentally and numerically for obtaining a more homogeneous ow distribution and thus reducing fouling. CFD numerical simulations were performed using CFX (ANSYS, Canonsburg, USA). The geometry of the channel was modeled with grid of 45,000 elements and the k turbulence model was used. The plate wall temperature was specied as boundary condition and ow and heat

7134

Flavio C.C. Galeazzo et al. / Chemical Engineering Science 61 (2006) 7133 7138
Table 1 Main dimensions of the FT-43 PHE
Dimension Plate length, between port centers (mm) Plate width, between gaskets (mm) Plate thickness (mm) Mean channel gap (mm) Equivalent diameter (mm) Port diameter (mm) Heat transfer area (m2 /plate) Value 90 60 1.0 1.5 3.0 6.5 0.0050

transfer were simulated for incompressible ow under steadystate conditions. The CFD predictions were in reasonable agreement with the experimental results of ow visualization and fouling, but discrepancies exist on the location of the recirculation zones. Grijspeerdt et al. (2003) carried out 2D and 3D CFD simulations of ow between two corrugated plates for investigating ow distribution. The grid contained 551,265 points, the wall temperature was specied, the BaldwinLomax turbulent stress model was used and the software FINE-Turbo was employed (NUMECA, Brussels, Belgium). The authors point that CFD could be a valuable tool for optimizing the design of PHEs for milk processing with respect to fouling. However, the model would need to simulate the dynamic protein denaturation and adhesion (Georgiadis and Macchietto, 2000). Moreover, the tested geometry differs from that of a PHE channel where the plates have opposing corrugation orientations that provide a series of contact points between plates. Fernandes et al. (2005) used the CFD tool POLYFLOW (FLUENT, Lebanon, USA) to model stirred yogurt ow between two corrugated plates. A geometrical domain with 173,634 elements represented half of the corrugated core of a PHE chevron channel since symmetry was assumed for simplication. Non-isothermal ow with non-Newtonian behavior (HerschelBulkley rheological model) was considered. The geometrical domain was highly complex due to the multiple contractions and expansions along the channel and the mesh was constituted by tetrahedral, hexahedral and pyramidal elements. CFD predictions and experimental data were in very good agreement with a mean error of 7% for outlet temperatures. For this paper, a CFD model was used to simulate ow and heat transfer in a four-channel PHE with at plates and two distinct congurations (parallel and series ow). An accurate geometric representation of the PHE was generated, including channels, plates and conduits, for creating a virtual prototype of the exchanger that takes into account both unequal ow distribution among channels and non-uniform ow distribution inside the channel. Moreover, the heat transfer between channels was modeled instead of specifying an estimated wall temperature as boundary condition. Experimental results were compared to numerical predictions from the 3D CFD model and also from a 1D plug-ow model for PHEs with generalized congurations. 2. Experimental procedure

ter from a thermostatic bath with temperature control (temperatures tested: 50, 70 and 80 C) in a closed circuit with a centrifugal pump (ow rates tested: 0.6, 0.8 and 1.0 L/min). Cold water was obtained from a closed circuit connected to a chiller unit (ow rates tested: 0.6, 0.8 and 1.0 L/min). Two distinct pass arrangements were used: parallel 12/12 and series 2 1/2 1, both countercurrent. The plate pack was mounted between two large plastic connector plates for minimizing heat losses to the ambient. Inlet and outlet stream temperatures were continuously monitored with calibrated thermocouples connected to a personal computer and volumetric ow rates were monitored with calibrated ow meters for the closed circuits and by direct measurement for the water from the inlet tank. In each run, after steady-state conditions were veried, temperatures were recorded for at least 2 min and the mean and standard deviation were calculated. Different process conditions were obtained by varying temperature and ow rate of the inlets. Runs were accepted only if the deviation between heat loads from hot and cold sides were less than 15% and the standard deviation for the stream temperatures were less than 1.0 C. Eqs. (1a) and (1b) were used for determining the heat load for the hot and cold sides of the PHE, where W is the mass ow rate and Cp is the specic heat of water for the mean stream temperature. The experimental heat load is obtained from the mean between Qhot and Qcold . Qhot = Whot Cp,hot (Thot,in Thot,out ), Qcold = Wcold Cp,cold Tcold,out Tcold,in . 3. Mathematical modeling 3.1. 3D CFD modeling (1a) (1b)

The study was conducted using an Armeld FT-43 laboratory plate pasteurizer (ARMFIELD, Hampshire, UK), which is described with more details by Gut et al. (2004). The PHE of the FT-43 has 12 8 cm stainless steel at plates with silicon gaskets; the main dimensions are presented in Table 1. Test uid was distilled water on hot and cold sides of the PHE. For varying temperature and ow rate of the hot and cold streams, the product, heating and cooling lines of the pasteurizer were used for the experimental runs. The product line provided water at ambient temperature using a peristaltic pump (ow rates between 0.3 and 0.8 L/min). The heating line provided hot wa-

The rst step of the CFD simulation was the mesh generation, which is the geometrical domain where the equations for heat and momentum transfer are solved (see Fig. 1). The models contain approximately one million hexahedral elements and were created with the software GAMBIT (FLUENT Inc., Lebanon, USA). Fig. 2 shows the expanded CFD representation of the two PHE congurations without the plates. The second step was the establishment of the boundary conditions and material properties. The hot and cold inlet boundary conditions were set as velocity inlets, with the corresponding

Flavio C.C. Galeazzo et al. / Chemical Engineering Science 61 (2006) 7133 7138

7135

Fig. 1. (a) The plate pack of the PHE used in this study. (b) The CFD representation of the uid domain for series ow arrangement showing a mesh detail.

Tests were also conducted to verify the grid independency of the results. Simulations were performed with decreasing size grids, until the results were consistent. The variable used in this comparison was the mean outlet temperature. The problem was numerically solved using the nite volume method with the software FLUENT 6.1.22 (FLUENT, Lebanon, USA). The simulations were carried in a Pentium 4 workstation with 2 GB RAM. Each simulation took approximately 5 h to converge, e.g. all equations reached a scaled residual value of 1 104 and the temperature monitor on the hot outlet was stable. The predicted heat load is determined from Eqs. (1a) and (1b) using the outlet temperatures obtained from the CFD simulation. 3.2. 1D plug-ow model In this work, experimental results were also compared to predictions of heat load obtained from the 1D plug-ow model presented by Gut and Pinto (2003), which determines the temperature distribution throughout the PHE channels and makes use of typical NusseltReynoldsPrandtl correlations for the determination of the mean convective heat transfer coefcients. The heat transfer parameters of the model were adjusted to t experimental data collected from 121 runs using various congurations of the FT-43 PHE with different numbers of plates (Gut et al., 2004). It was veried that the thermal model is well adjusted using data collected from a single PHE conguration. However, when tting the model using data collected from various congurations, considerable differences were obtained between experimental and predicted heat loads. These results show that the velocity and heat transfer coefcient distributions inside the PHE must be dependent on its conguration, and a single heat transfer correlation may not adequately represent all possible congurations of a PHE. The adjusted correlation, Nu = 0.457Re0.771 Pr0.333 (R 2 = 0.82), was used in this work to determine the mean convective heat transfer coefcients inside the PHE. The 1D model was used for obtaining the outlet temperatures and predicted heat load is determined from Eqs. (1a) and (1b). 4. Results and discussion The numbers of valid experimental runs were 37 and 32 for the series (2 1/2 1) and parallel (1 2/1 2) arrangements, respectively, with varying ow rate and inlet stream temperature. Table 2 presents the range of the main variables for the experimental data. Experimental and simulation results obtained for heat load are shown in Figs. 3 and 4. For the series arrangement, the mean error for heat load prediction was only 8% for both 3D and 1D models. However, for the parallel arrangement, the mean error was 25% for 1D model and 12% for the 3D model. Since the series arrangement does not have the problem of ow maldistribution among channels, as in the parallel arrangement, a better agreement with experimental data was expected for the series arrangement.

Fig. 2. Expanded CFD representation and opened plate pack for (a) parallel arrangement (1 2/1 2) and (b) series arrangement (2 1/2 1).

ow rates and the temperatures according to the experimental data, and the outlets were set as pressure outlets. The metal plates were modeled as thin walls with the thermal resistance of a 1 mm stainless steel wall. All the exterior walls were modeled as adiabatic. The uid domains were modeled with the properties of water. The temperature correlations for density, specic heat, viscosity and thermal conductivity used in all simulations are presented elsewhere (Gut and Pinto, 2003). The sensitivity of the simulation to the turbulence modeling was investigated. The simulation was solved using the laminar ow model, the k turbulence model with wall functions and the k turbulence model with enhanced wall treatment. It was veried that the ow inside the exchanger was mostly laminar; however, there were turbulent regions, especially near the plate ports. In order to correctly model the ow and heat transfer, the k turbulence model with enhanced wall treatment model was chosen, which means that the k model was solved to the wall with no aid of length-scale-based wall functions. Moreover, the experimental results for outlet temperature were in better agreement with the simulation using this model.

7136
Table 2 Ranges of experimental data

Flavio C.C. Galeazzo et al. / Chemical Engineering Science 61 (2006) 7133 7138

Series (2 1/2 1) Min. Hot side W (kg/h) Tin ( C) Tout ( C) v (m/s) Re W (kg/h) Tin ( C) Tout ( C) v (m/s) Re Q (W) 16.2 13.2 11.4 5.0 136 12.1 3.5 5.0 3.7 152 70 Max. 73.6 80.9 74.1 22.7 1528 63.0 36.6 54.1 19.4 894 749

Parallel (1 2/1 2) Min. 36.3 50.1 45.4 5.6 298 15.5 5.6 11.4 2.4 93 179 Max. 69.2 80.9 74.1 10.7 764 61.8 28.5 48.0 9.5 324 664

Cold side

Exchanger

800 3D (CFD) 1D (Plug-flow) Predicted Heat Load (W) 600 - 15% + 15%

800 3D (CFD) 1D (Plug-flow) Predicted Heat Load (W) 600 - 15% + 15%

400

400

200

200

0 0 200 400 600 800 Experimental Heat Load (W)

0 0 200 400 600 800 Experimental Heat Load (W)

Fig. 3. Experimental and simulation results for heat load: series arrangement (2 1/2 1).

Fig. 4. Experimental and simulation results for heat load: parallel arrangement (1 2/1 2).

It can also be observed in Figs. 3 and 4 that the heat load predicted with the CFD model is consistently higher than the experimental value. Since the CFD model does not account for ambient heat losses, this behavior was expected. For the 1D model, it can be observed that most of the points are below the 45 line in Figs. 3 and 4, e.g. the predicted heat load is lower than experimental value. This systematic error is originated from unrealistic assumption of plug-ow of the PHE model for generalized congurations. The 1D model was adjusted to t experimental data collected from various congurations. Though the overall model adjustment was satisfactory, as shown by Gut et al. (2004), some congurations had a generally positive deviation (predicted heat load higher than experimental) and others had a generally negative deviation (predicted heat load lower than experimental), as those studied in this work (2 1/2 1 and 1 2/1 2). The prediction errors for heat load were also correlated with C (ratio of heat capacities, W Cp , of hot and cold sides,

with C 1.0) and with the thermal effectiveness of the exchanger = Q/Qmax , where Qmax is the maximum allowable heat load. It was veried that the errors of the 1D model are clearly dependent on the operational conditions (high C and low result on low errors), whereas the errors of the 3D are not. When using a simulation model, it is highly desired that the prediction error is not dependent on the operational conditions. The detailed results from the CFD model allow the analysis of velocity and temperature distribution inside the PHE. Hightemperature regions and stagnation areas could be observed, indicating regions susceptible to fouling. An example of temperature and velocity distribution is presented in Figs. 5 and 6 for a series arrangement (cold side inlet: 61.3 kg/h at 80.5 C; hot side inlet: 61.4 kg/h at 36.5 C). Movie 1 (supplementary multimedia data) combines the velocity and temperature distributions for the cold and hot sides of the PHE for a better view of the CFD simulation results.

Flavio C.C. Galeazzo et al. / Chemical Engineering Science 61 (2006) 7133 7138

7137

Fig. 5. Midplane temperature distribution for the channels of a series ow arrangement: 3D and 1D model results.

Fig. 6. Midplane velocity distribution for the channels of a series ow arrangement: 3D and 1D model results.

The results obtained from the 1D plug-ow model are also included in Figs. 5 and 6. Note that the 1D model considers only the plate length between ports for heat transfer and that a uniform velocity distribution throughout the channel is assumed. It was not possible to obtain experimental results of temperature and velocity distribution inside the PHE for comparison with the simulation results. Flow visualization techniques, such as used by Cave et al. (1983), Kho and Mller-Steinhagen (1999) and Grijspeerdt et al. (2003), would be required. Although the errors for heat load prediction through the 1D model were close to those obtained through CFD, the 1D model has the major disadvantage of requiring a large number of experimental runs using various congurations for adjusting the

heat transfer and pressure drop correlations. On the other hand, the 3D CFD model has a limitation on the computational time required for model simulation. However, with the development of faster computer processors it will be possible to simulate more complex PHE geometries, with higher number of channels or running in transitory mode. 5. Conclusions Using a CFD tool it was possible to build a virtual prototype of a PHE with four channels and at plates. The simulation results included outlet temperatures, heat load, as well as the 3D temperature and velocity distribution. For the series

7138

Flavio C.C. Galeazzo et al. / Chemical Engineering Science 61 (2006) 7133 7138

arrangement the prediction errors for heat load obtained from the CFD and the plug-ow model were similar (8%); however, the CFD results for parallel ow were in better agreement with experimental data. A low dependence of the prediction errors was also veried when using the CFD model, in comparison with the plug-ow model. The main advantages of the CFD model of the PHE are the detailed temperature and velocity distributions obtained and the fact that it is not necessary to collect extensive experimental data to adjust the model parameters. On the other hand, the computational time required is a limitation for modeling PHEs with a large number of plates or more complex geometries using CFD. Notation AC C Cp De h k N Nu Pr Q Re T v W cross-sectional area for channel ow, m2 heat capacities ratio, C 1.0, dimensionless specic heat at constant pressure, J/kg C equivalent diameter of the channel, m convective heat transfer coefcient, W/m2 C thermal conductivity, W/m C number of channels per pass Nusselt number, Nu = h De /k , dimensionless Prandtl number, Pr = Cp /k , dimensionless heat load, W Reynolds number, Re = W De /( N AC ), dimensionless temperature, C velocity, m/s mass ow rate, kg/s

Appendix A. Supplementary data Supplementary data associated with this article can be found in the online version at 10.1016/j.ces.2006.07.029. References
Bassiouny, M.K., Martin, H., 1984. Flow distribution and pressure drop in plate heat exchangersI, U-type arrangement. Chemical Engineering Science 39 (4), 693700. Cave, G.D., Giudici, M., Pedrocchi, E., Pesce, G., 1983. Study of uid ow distribution inside plate heat exchangers by thermographic analysis. In: Taborek, J., Hewitt, G.F., Afgan, N. (Eds.), Heat Exchangers Theory and Practice. McGraw-Hill, New york, pp. 521531. Fernandes, C.S., Dias, R., Nbrega, J.M., Afonso, I.M., Melo, L.F., Maia, J.M., 2005. Simulation of stirred yoghurt processing in plate heat exchangers. Journal of Food Engineering 69, 281290. Georgiadis, M.C., Macchietto, S., 2000. Dynamic modelling and simulation of plate heat exchangers under milk fouling. Chemical Engineering Science 55, 16051619. Grijspeerdt, K., Hazarika, B., Vucinic, D., 2003. Application of computational uid dynamics to model the hydrodynamics of plate heat exchangers for milk processing. Journal of Food Engineering 57, 237242. Gut, J.A.W., Pinto, J.M., 2003. Modeling of plate heat exchangers with generalized congurations. International Journal of Heat and Mass Transfer 46 (14), 25712585. Gut, J.A.W., Pinto, J.M., 2004. Optimal conguration design for plate heat exchangers. International Journal of Heat and Mass Transfer 47, 48334848. Gut, J.A.W., Fernandes, R., Pinto, J.M., Tadini, C.C., 2004. Thermal model validation of plate heat exchangers with generalized congurations. Chemical Engineering Science 56, 45914600. Kho, T., Mller-Steinhagen, H., 1999. An experimental and numerical investigation of heat transfer fouling and uid ow in at plate heat exchangers. Transactions of the IChemE 77A, 124130. Mehrabian, M.A., Poulter, R., Quarini, G.L., 2000. Hydrodynamic and thermal characteristics of corrugated channels: experimental approach. Experimental Heat Transfer 13, 223234. Rao, B.P., Kumar, P.K., Das, S.K., 2002. Effect of ow distribution to the channels on the thermal performance of a plate heat exchanger. Chemical Engineering and Processing 41, 4958. Tereda, F.A., Srihari, N., Das, S.K., Sunden, B., 2005. Experimental study on port to channel ow distribution of plate heat exchangers. In: Shah, R.K., Ishizuta, M., Rudy, T.M., Wadekar, V.V. (Eds.), Proceedings of the Fifth International Conference on Enhanced, Compact and Ultra-Compact Heat Exchangers, pp. 208214. Thonon, B., Mercier, P., 1996. Les changeurs plaques: dix ans de recherche au GRETh: partie 2. dimensionnement et mauvaise distribuition. Revue Gnrale de Thermique 35, 561568. Xia, B., Sun, D.-W., 2002. Applications of computational uid dynamics (CFD) in the food industry: a review. Computers and Eletronics in Agriculture 34, 524. Zaleski, T., 1984. A general mathematical-model of parallel-ow, multichannel heat-exchangers and analysis of its properties. Chemical Engineering Science 39 (7/8), 12511260.

Greek letters thermal effectiveness, % uid viscosity, Pa s Subscripts cold hot in max out cold uid hot uid uid inlet maximum uid outlet

Acknowledgments The authors would like to acknowledge nancial support from FAPESP (The State of So Paulo Research Foundation).

Das könnte Ihnen auch gefallen